• Nem Talált Eredményt

This manuscript is contextually identical with the following published paper:

N/A
N/A
Protected

Academic year: 2022

Ossza meg "This manuscript is contextually identical with the following published paper:"

Copied!
41
0
0

Teljes szövegt

(1)

1 This manuscript is contextually identical with the following published paper:

1

Balogh Csilla; Vláčilová Alena; G.‐Tóth László; Serfőző Zoltán (2018) Dreissenid 2

colonization during the initial invasion of the quagga mussel in the largest Central 3

European shallow lake, Lake Balaton, Hungary. - JOURNAL OF GREAT LAKES 4

RESEARCH, 44: pp. 114-125.

5

The original published PDF available in this website:

6

https://www.sciencedirect.com/science/article/pii/S0380133017301910?via%3Dihub 7

8 9

Dreissenid colonization during the initial invasion of the quagga mussel in the largest 10

Central European shallow lake, Lake Balaton, Hungary 11

12

Csilla Balogha,b, Alena Vláčilovác, László G.-Tótha,b, Zoltán Serfőzőa,b*

13 14

aCentre for Ecological Research, Balaton Limnological Institute, Hungarian Academy of 15

Sciences Tihany, Hungary 16

bMTA Centre for Ecological Research, GINOP Sustainable Ecosystems Group, 8237 Tihany, 17

Klebelsberg Kuno u. 3.

18

cCentre for Popularization, Faculty of Science, Palacky University, Olomouc, Czech Republic 19

20

*Corresponding author 21

address: MTA CER BLI, Klebelsberg Kuno street 3, Tihany, H-8237 22

mailto: serfozo.zoltan@okologia.mta.hu 23

phone: +3687448244/202 24

(2)

2 ABSTRACT

25 26

The colonization progress of the invasive bivalve dreissenids, the formerly dominant 27

Dreissena polymorpha and the recently (2008) introduced Dreissena rostriformis bugensis 28

was studied between 2009 and 2013 in the largest Central European shallow lake, Lake 29

Balaton, Hungary. The density of dreissenid planktonic veligers, new settlers (post-veligers 30

and early juveniles), and the population structure (density, length frequency, relative 31

abundance) of the two species were monitored on experimentally introduced natural stone 32

substrata, on different time scales. Dreissenids started dynamic settling following a sudden 33

veliger bloom. As substratum saturation progressed, competition between species for places 34

was suggested, which, after two years, led to an increased number of large individuals (> 20 35

mm) and also recruits of D. r. bugensis. By contrast, the population of D. polymorpha 36

confined to middle size (11-18 mm) individuals of the first settler generation. On local 37

substrata, where the benthic community was already established, the replacement of D.

38

polymorpha by D. r. bugensis took longer, but it happened in a similar way. The invasion 39

speed of D. r. bugensis in Lake Balaton resembled the speed obtained in other European 40

water bodies where D. r. bugensis, similar to Lake Balaton, was introduced much later than 41

D. polymorpha. However, a longer replacement process was found in North America, where 42

both species invaded new habitats at the same time. This suggests that the speed, and 43

probably the success, of D. r. bugensis invasion depends on new surface availability, and 44

whether the two dreissenid species are introduced together or at different times.

45 46

Index words: biological invasion, colonization dynamics, dreissenid, quagga mussel, 47

population structure 48

(3)

3 Introduction

49 50

Invasive aquatic alien species, among them the freshwater bivalve dreissenids 51

(Dreissena polymorpha, Pallas, 1771 [D. polymorpha, zebra mussel]) and Dreissena 52

rostriformis bugensis, Andrusov 1897 [D. r. bugensis, quagga mussel]) are serious threats to 53

and provoke dramatic changes in community abundance, species diversity, resource 54

availability, nutrient cycling and functioning of the ecosystem (Crooks, 1998; Gutiérrez et al., 55

2003; Ward and Ricciardi, 2007; Ricciardi and MacIsaac, 2011). Due to their invasive nature 56

and intensive filter-feeding behavior, dreissenids are counted among the most invasive and 57

destructive invaders in North American and European freshwater ecosystems (OTA, 1993;

58

DAISE, 2003; Pimentel et al., 2005).

59

Due to the fast spread of invasive species nowadays, a new invader is more likely to 60

come into contact with a former dominant invader (van den Brink et al., 1993; van der Velde 61

et al., 1994; Dick and Platvoet, 2000; Platvoet et al., 2009; Ricciardi, 2001; Gallardo and 62

Aldridge, 2015; de Gelder et al., 2016). Although the occurrence and progression of D.

63

polymorpha in Europe has long been described in many different habitats (Bij de Vaate, 1991;

64

Pollux et al., 2003; Aldridge et al., 2004; Cianfanelli et al., 2010; Karatayev et al., 2010a;

65

Palau Ibars et al., 2010; Stańczykowska et al., 2010), in contrast to North America (Mills et 66

al., 1993; Watkins et al., 2007; Dermott and Dow, 2008; Nalepa et al., 2010; Karatayev et al., 67

2013a), and to the Ponto-Caspian region where dreissenids originated (Orlova et al., 2004;

68

Zhulidov, et al., 2010), the characteristics of its population dynamics during the invasion of 69

the new dreissenid, D. r. bugensis, has only rarely been observed and investigated (Bij de 70

Vaate, 2010; Aldridge et al., 2014; Heiler et al., 2013; Matthews et al., 2014).

71

The dispersal speed of D. polymorpha was found to be much faster than that of D. r.

72

bugensis throughout their invasion history (Karatayev et al., 2011). Therefore, in Europe, 73

(4)

4 apart from the connected river systems (Orlova et al., 2004; Zhulidov et al. 2010; Heiler et al., 74

2012; Heiler et al., 2013; Matthews et al., 2014) which are known as invasion pathways, in 75

lakes and reservoirs, D. r. bugensis has only appeared in the Kuybyshev, Saratov and Rybinsk 76

Reservoirs (part of the Ponto-Caspian Volga River invasion corridor, Orlova, 2004), some 77

Dutch lakes (Lake Ijsselmeer and Markermeer, Bij de Vaate et al., 2013), and recently in 78

Great Britain (Aldridge et al., 2014). In contrast to the spread over long distances, in 79

European water bodies where D. r. bugensis was introduced, its population increase rate was 80

26% per year within the dreissenid community, until the displacement of D. polymorpha 81

(Heiler et al., 2013), which suggests the significant spread of this species in local habitats.

82

Similar trends were reported from the Laurentian Great Lakes, where five years after the 83

initial introduction, D. r. bugensis represented 44% of the dreissenid population (Dermott and 84

Dow, 2008), and later increased up to 97% (Patterson et al., 2005), leading to total 85

replacement of D. polymorpha in the deep parts of the lakes, and the balanced co-existence of 86

the two dreissenids in shallow parts (Patterson et al., 2005; Watkins et al., 2007; Nalepa et al.

87

2010). It was also found that in some shallow lakes and reservoirs, or in shallow areas of deep 88

lakes, dreissenids could live together for longer periods of time in the same habitat (Tzeyeb et 89

al.1966; Zhulidov et al., 2010; Karatayev et al., 2011, 2013b), or exhibit a reversal, where D.

90

r. bugensis was displaced by D. polymorpha (Zhulidov et al., 2006). Thus, the interaction 91

between the two dreissenids can be characterized as competitive niche partitioning; however, 92

the dynamics and progression of the co-existence remain unknown, and they seem to be 93

influenced by environmental conditions.

94

The site of the present study, Lake Balaton, is the largest shallow lake in Central Europe 95

connected to the Danube River via the Sió channel (Fig. 1). In 1932, D. polymorpha was the 96

first Ponto-Caspian invader introduced from the Danube River to the lake, possibly via ship 97

transport (Sebestyén, 1938). Around 75 years later (2008), D. r. bugensis was discovered as a 98

(5)

5 new dreissenid invader in Lake Balaton (Majoros, 2009; Balogh and G.-Tóth, 2009; Benkő 99

Kiss, Á. personal communication). D. r. bugensis may have been imported in the same way as 100

D. polymorpha since the mussel was detected in the Hungarian region of the Danube River 101

much earlier than it was estimated to appear in Lake Balaton (Szekeres et al., 2008). The 102

recent introduction of D. r. bugensis to Lake Balaton makes this lake ideal to follow the 103

invasion dynamics in real time, and study the early consequences of the invasion on the 104

benthic community, including the dominant D. polymorpha population.

105

Studying the settlement and growth dynamics of the dreissenids would provide unique 106

data to help understand the population trends and invasion success of D. r. bugensis, thus 107

supporting the assessment and prediction of its environmental impact, ecological changes, and 108

management (Wong and Gerstenberg, 2011). Shallow lakes invaded by D. r. bugensis have 109

been reported by Karatayev et al. (2013b), who expected invasion dynamics to be different in 110

shallow versus deep lakes. For these reasons, we investigated the colonization progress of 111

dreissenids in Lake Balaton during the early invasion of D. r. bugensis. Colonization was 112

evaluated on early, short-term (daily), medium-term (monthly), and long-term (yearly) time 113

scales on experimentally introduced natural stone substrata by following the progression of 114

the number (relative abundance, density) and growth (size and covered substratum area) of 115

the two dreissenids. Data were compared with those sampled from the local stone substrata.

116

Such a complex research approach, providing high resolution data in a medium-term study, 117

has not yet been used for studying the invasion of D. r. bugensis. The objective was to 118

analyze the colonization dynamics of dreissenids, in the knowledge that D. r. bugensis has 119

been introduced to the lake recently, and the environmental conditions favor its settlement.

120 121

Materials and Methods 122

123

(6)

6 Study site with local conditions

124 125

The support structure for experimental substrata submerged in Lake Balaton was placed 126

near to the riprap zone at 1.9 m depth, in front of the shoreline of the Balaton Limnological 127

Institute, Tihany (Fig. 1). To compare the data obtained from the experimental substrata with 128

natural trends, three sampling locations (T1, T2, T3) in the riprap along the shoreline of the 129

Tihany peninsula were selected. The water characteristics of these local bottom substrata 130

(type, size, shape, depth of their location [T1: 1.5 m; T2: 1.9 m; T3: 1.1 m]) were similar to 131

those used for the experimental substrata. The only difference was that the local substrata had 132

already been coated by natural biofilm, including dreissenids.

133

During each substratum sampling, at the same depth where the support structure for 134

experimental substrata was placed, and the local sampling sites were situated, beside the 135

substrata, water characteristic variables were recorded in the conventional manner, using a 136

Horiba U-10 water multiparameter measuring instrument. In the first few months of the study, 137

between the summer and winter period, the measured values were as follows: temperature: 7–

138

29 °C; conductivity: 751–910 µS cm-1; pH: 8.0–8.9; chlorophyll-a concentration: 5.4±3.15 μg 139

L-1, the highest content of suspended material: 25 mg (dw) L-1. 140

141

Support structure for experimental substrata 142

143

To study the progression of surface colonization by dreissenids, stone holders were 144

created from a massive metal frame measuring 1×1 m in size, and joined with protruding 145

pipes, distributed at equal distances, and serving as holders (Fig. 2). Sixty palm-sized (surface 146

was 0.0191±0.0038 m2/ each) red sand stones of similarly irregular shape (hereafter referred 147

to as ‘stones’) were collected from the shoreline, scrubbed and dried. They were drilled, 148

(7)

7 inserted, and fixed onto metal pipes of uniform size with commercial glue. Stones with pipes 149

could be slipped into and easily taken out of the holder pipes, ensuring a secure fit on the 150

frame and sampling.

151 152

Sample collection 153

154

In each sampling session, three stones were randomly selected for retrieval by diving 155

from both natural and experimental substrata. Each stone was placed in a plastic bag for 156

protection, and transported to the laboratory, where the encrustation of the substrata was 157

removed by knife and soft brush, and sieved through a 60 m nylon net. The collected animal 158

specimens were preserved in 70% ethanol.

159

Veligers were sampled with a 50 cm high (volume: 34 L) Schindler–Patalas sampler 160

equipped with a 60 m mesh-sized collector funnel near to the site, where experimental 161

substrata were placed, on each occasion the substrata was sampled. Sampling was carried out 162

at 50 cm increments along the entire vertical depth. The samples were pooled together and 163

concentrated to 20 mL, then preserved with ethanol.

164 165

Sampling plan (dates and terms) and weather characteristics 166

167

The study began by submerging the instrument holding the experimental substrata on 3 168

August 2009. Sampling started from the second day of submersion and continued at a three- 169

day frequency in August (short-term scale), monthly from September to December of 2009 170

(medium-term), and annually, beginning in the summer of 2010 for the following three years 171

(long-term). In parallel, the local substrata were examined each August during the study 172

period, until 2013.

173

(8)

8 The average summer temperature was 0.6 and 1 °C higher in 2012 than in 2010 and 174

2011, respectively. The hottest water temperature (29.1 °C) was also recorded in 2012, when 175

a long dry period caused a significant decrease in water levels, which was low, as parts of the 176

shoreline had dried up. In summer 2012, a number of days with stormy winds and waves were 177

registered, which created harsh turbulences in the shallow water. Extreme weather 178

phenomena, such as heavy storms (wind speed was 80–90 km/h) on the 5th and the 14th of 179

August, and rapid cooling thereafter in September, occurred.

180 181

Density and size measurements 182

183

To calculate the surface area of sampled stones, the entire surface of the stones were 184

traced onto wrapping paper. An algorithm for cut paper weight vs. surface area was derived.

185

The density of dreissenids was represented as ind m-2 stone surface. The relative abundance of 186

D. r. bugensis within the whole dreissenid population was calculated and given as percentage 187

contribution in both introduced and local substrata.

188

The two dreissenids were distinguished from each other according to their unique 189

morphological features (Spidle, et al., 1995; van der Velde et al., 2010). The two species were 190

only differentiated if the dreissenid individual was >2 mm. New settlers were separated based 191

on their sizes: <0.5 mm (post-veligers – plantigrade) and 0.5< and <2 mm (early juveniles – 192

siphon-forming stage), according to Claudi and Mackie (1994), Kirpichenko (1964), and 193

Ackerman, et al. (1994).

194

The number of dreissenids >2 mm was counted with the naked eye, and their length was 195

measured with a digital caliper. Post-veligers and early juveniles were counted and measured 196

under a stereomicroscope using a mm scale underneath the counter dish. Length frequency 197

histograms were generated using 1 mm size classes to assess population size structure (Mills 198

(9)

9 et al., 1993; Orlova et al., 2004, 2013; Dermott and Dow, 2008; Karatayev et al., 2013a). To 199

count veligers, 2–5 mL multiple subsamples were taken from the concentrated 20 mL sample, 200

and examined with a Zeiss–Opton inverted microscope.

201 202

Calculation of substratum surface saturation by dreissenids 203

204

Cover of experimental substrata by dreissenids was calculated using the following 205

formula:

206

∑ length × [a × length + b] × π × n

(5−27) 4 Where,

207

length is the length of the individual in mm accuracy on a 5–27 mm scale;

208

a is the slope of the length–width regression line;

209

b is the intercept of the length–width regression line;

210

a × length + b is the calculated width;

211

length × [a × length + b] × π/4 is the surface of the ventral side of the animal;

212

n is the number of animals of a specific length;

213

Σ(527) is the total ventral surface of animals covering the surface.

214

The length– width correlation was obtained from the measurement of 30–40 D. polymorpha 215

and D. r. bugensis individuals, respectively, in each length group (between 5 and 27 mm with 216

1 mm difference). The longest distance of the rostro-caudal axis was measured as the length, 217

and the width was measured as the longest dimension in the direction perpendicular to the 218

length. Dreissenids face the attaching surface with their ventral side, which looks like an 219

ellipsoid in planar projection. The ventral side area of all mussels forming the dreissenid 220

population was therefore calculated using the formula for elliptical area (a × b × π, where a 221

and b are half of the length and width values, respectively), and finally, extrapolated to the 222

(10)

10 number of animals on the substrata. The sum of the surface area of differently sized 223

dreissenids colonizing the substrata simultaneously gave a total surface area, which would 224

ideally cover the substrata if the animals settled side by side. This score overestimates the real 225

surface area occupied by dreissenids, since animals attach to the substrata by their byssus and 226

thus do not occupy the surface with their total base (ventral surface). In addition, as 227

colonization progresses, many of the animals also use each other’s shells as settling sites 228

(multilayer aggregation). Nevertheless, with knowledge of these shortcomings, this 229

calculation allows to assess the dynamics of substratum occupation and estimate the time 230

when dreissenids reach total occupation.

231 232

Statistics 233

234

Before analysis, datasets were transformed to achieve homogeneity of variance and 235

improve normality. The normality of the data was checked with a normal Q-Q plot of the 236

model residuals (Sokal and Rohlf, 1995). A mixed model ANOVA was used to analyze 237

differences in density (log-transformed) and average length (log-transformed) between the 238

two species, on different time scales. The studied variables were species (within-subject, 239

repeated measures factor) and time (between-subject factor). Separate analyses were carried 240

out for different time scales (months, years). Sequential Bonferroni-corrected t-tests were 241

used to show the significant interactions.

242

Yearly differences in the relative abundance of D. r. bugensis (percentage of D. r.

243

bugensis in the dreissenid population) were compared between sampling sites using GLM 244

ANOVA. The relative abundance of D. r. bugensis as a dependent (log-transformed) variable 245

was analyzed with the independent categorical variables (fixed factors), time and sites 246

(including experimental and local substrata) in the model. A Tukey test was used as a post- 247

(11)

11 hoc procedure for evaluating the main effects and interactions (for differences between the 248

sites at a certain date, and differences between years at a given site).

249 250

Results 251

252

Densities of planktonic veligers and new settlers (size: 0.5–2 mm) on experimental substrata 253

254

Within a week of the onset of the settlement study, even though the veliger 255

concentration in the water was low (1-1.3 ind L-1), early colonization of post-veligers was 256

already observed (Fig. 3). Shortly thereafter, an extraordinary boom of veliger expansion was 257

detected (on the 14th August, 2009; 376±74 ind L-1), which was limited only to single day, 258

and could not be seen for the remainder of the month. A week after the veliger boom, on the 259

23rd of August, a peak appeared in the density of post-veligers (11500±6100 ind m-2) which, 260

within a month, showed a downward trend to a similar level as observed before the boom.

261

The curve of density dynamics of early juveniles was similarly shaped, peaking in September 262

with 27000±12000 ind m-2. As the temperature dropped, and autumn transitioned to winter, 263

veligers disappeared from the water sample, and post-veligers and early juveniles remained at 264

low density levels on the substratum surface (2009 December, post-veligers: 300±130 ind m-2; 265

early juveniles: 2600±300 ind m-2). In subsequent summers, veligers were consistently found 266

at moderate levels (14-25 ind L-1), whereas post-veligers and early juveniles showed a 267

decreasing tendency to colonize with a density between 400–1400 ind m-2,and 3400–11100 268

ind m-2, respectively.

269 270

Dreissenid (size: >2 mm) density on experimental substrata 271

272

(12)

12 In the following months after substratum deployment (August 2009), no significant 273

differences in density data were found either between the two dreissenids, or between the 274

months (Fig. 4, Table 1a). The density of each species changed during this period, between 275

2652 and 2699 ind m-2. From 2010 to the end of the study (2012), the density of D. r.

276

bugensis significantly increased year by year (Fig. 4, Table 1b.), in contrast to D. polymorpha 277

for which density slightly fluctuated, but did not change overall. By 2012, D. r. bugensis 278

density reached up to 42453±10321 ind m-2, which was six times higher than that of D.

279

polymorpha.

280 281

Relative abundance of dreissenids (size: >2 mm) on experimental and local substrata 282

283

No difference in relative abundance was found on experimental substrata between the 284

two dreissenids after one year of substratum implantation (Fig. 5). From 2010 onward, with 285

increasing differences year by year, the abundance of D. r. bugensis (percentage of D. r.

286

bugensis in the dreissenid community) significantly exceeded that of D. polymorpha, which 287

was accompanied by the decline of the latter population (Fig. 5, Table 3).

288

Along with sampling from the experimental substrata during the study period (2009- 289

2012), in each August, and also in the upcoming year (2013), the dreissenid population was 290

also examined on natural (local) substrata at three points of the Tihany peninsula (see the 291

sampling site map in Fig. 1). In the year of substratum implantation, the relative abundance of 292

D. r. bugensis was significantly different at all three points (Fig. 5, Table 3). Percentage of D.

293

r. bugensis was 29.6±8% in T1, 48.6±7.1% in T2, and 16.4±4.4% in T3. By 2010, the 294

percentage of D. r. bugensis had significantly increased in all sampling sites, and a difference 295

was only found between the T2 and T3 sites. At the same time, except in the T2 site, relative 296

abundance was similar in the introduced as well as the local substrata. In 2011, the percentage 297

(13)

13 of D. r. bugensis was further increased in experimental substrata, and T1. Trends of relative 298

abundance equalization continued between the sampling sites. This resulted in around 80%

299

relative abundance of D. r. bugensis in all sites and no annual differences between the sites by 300

2012. The decrease in water level resulted in high mortality for the dreissenid populations on 301

the riprap, along which piles of shells could be traced in 2012; however, this had no 302

significant impact on the growing relative abundance of D. r. bugensis, which reached 98%

303

by 2013.

304 305

Size related composition of dreissenids (size: >2 mm) on experimental substrata 306

307

Fine resolution analysis of length distribution showed that both species were equally 308

represented on the surface after two months of substratum deployment (Fig. 6). The most 309

common sizes were within the 2–9 mm range, forming a bell shaped distribution.

310

Interestingly, some adult animals also appeared on the substrata at this very early stage of 311

colonization. In the subsequent months, the distribution of individuals within the 3–13 mm 312

range equalized further, which resulted in a smoother distribution pattern (Fig. 6). By the end 313

of 2009, five months after colonization started, the largest animals (> 12 mm) were mainly D.

314

r. bugensis. A year later (2010), and in 2011, it was again evident that D. r. bugensis was 315

more frequent among the largest animals (>12 mm), whereas the frequency of adult animals 316

belonging to the size range of 10–13 mm was higher in D. polymorpha. Following the annual 317

changes in length during the examination period, the whole size distribution pattern of D. r.

318

bugensis (ranging from 2 to 28 mm) was found to be much wider than that of D. polymorpha 319

(ranging from 2 to 17 mm). In 2010, the most abundant sizes found in the D. polymorpha 320

population were within a narrow range of 7–11 mm in length, which slowly shifted to 10–16 321

mm by 2012. In contrast, the size frequencies of D. r. bugensis showed a rather heterogeneous 322

(14)

14 distribution, resulting in a less coherent size distribution in the plot. In 2010 and 2011, 323

dreissenids 4–7 mm in size, were missing or underrepresented in the samples.

324

The average length of settled D. r. bugensis reached its maximum as early as in 2010, 325

whereas that of D. polymorpha in 2011-2012 (Fig. 7). As a straightforward consequence of 326

the difference in the size distribution observed between the two dreissenids (see Fig. 6), the 327

average length showed significant differences from December 2009 until 2011 (Fig. 7, Table 328

2a, 2b). By contrast, a difference in the average length was not seen in 2012.

329 330

Saturation of dreissenids (size: >2 mm) on experimental substrata 331

332

Both dreissenid species showed similar, linear correlations between length and width 333

(Fig. 8a, b). The total surface occupied by the dreissenid population was estimated according 334

to the ventral (attaching) surface of individuals and the quantity of settled animals of different 335

sizes (see Fig. 6). Two months after deployment of the support structure for the experimental 336

substrata, the total surface of colonizing dressenids occupied 10% of the available surface of 337

the implanted stones (Fig. 8c). In the following cold season, this area did not increase 338

significantly until the end of 2009. A year after deployment, the number of dreissenids 339

attached to the substrata represented 122±28 cm2 surface, which, supposing idealistic and 340

homogeneous distribution, covered the whole available surface (Fig. 8c). In the forthcoming 341

years (2011, 2012), the total surface of settled animals slightly increased, exceeding that of 342

the substrata. Considering that surface occupation was overestimated, dreissenids could have 343

saturated the whole surface around the summer of 2011. However, empty spots on 344

experimental substrata could be seen until 2012 (Fig. 8d). From that time, the stratified 345

appearance of dreissenid populations and phenomenon of multilayer aggregation could be 346

more frequently observed on the experimental substrata (Fig 8d).

347

(15)

15 348

Discussion 349

350

Dreissenid population structure and dynamics on experimental and local substrata 351

352

Shortly after the experimental substrata were submerged into the lake, a huge boom of 353

dreissenid larvae was observed, which could also be seen in the increasing number of post- 354

veliger and early juvenile individuals attached to the new surface in the subsequent week and 355

month. The spawning season of dreissenids lasts from late March to November depending on 356

the lake temperature. The frequency of dreissenid larvae release has not been studied so far in 357

Lake Balaton, but it is assumed to be influenced primarily by weather extremities, which 358

occur more often nowadays rapidly changing the physical conditions of the shallow lake.

359

Simultaneous release of a large number of larvae within days in the middle of the spawning 360

period assumes the presence of some triggering substances, which affect gonadal activity and 361

promote sudden rather than smooth production of new larvae. The veliger release coincided 362

with a storm, which evoked big waves and mixed up the whole water column, increasing the 363

amount of the suspended material. Nevertheless, correlation between the storm and the veliger 364

release cannot be established, which was supported by the low number of veligers observed 365

also at a stormy day just after the settlement. Taking into account that D. polymorpha larvae 366

stay in the plankton for at least 7–15 days (Marsden, 1992; Ackerman et al., 1994), and 367

because of the irregular larvae release, this implies that to address dreissenid larvae 368

propagation, at least weekly sampling frequency is required.

369

It can be deduced from the density and length frequency data of dreissenids, which have 370

been taxonomically identified (> 2 mm), that D. r. bugensis and D. polymorpha colonized the 371

experimental substrata with equal success in the first few months. This theoretically suggests 372

(16)

16 that at the beginning of the colonization study, veligers were distributed equally between the 373

two species. However, due to its stronger attachment, D. polymorpha is a better colonizer and 374

more often remain on the substratum (Peyer et al., 2009, Collas et al., 2016), also suggesting 375

that D. polymorpha might have represented more on the substratum, albeit its veliger is less 376

abundant in the water.

377

Some large individuals, found unexpectedly in the samples of early colonization, might 378

come from the neighboring riprap by detaching and transporting via water currents or by 379

active locomotion to the experimental substrata.

380

After an equalized abundance, from the first year of colonization, the D. r. bugensis 381

population significantly increased on experimental and local substrata, implying the success 382

of this dreissenid over the other species in the colonization process. However, the progression 383

of D. r. bugensis colonization on local substrata took longer than on experimental substrata at 384

the beginning of the study. This might be because of the established benthic community on 385

local substrata including D. polymorpha and C. curvispinum, which saturated the surface and 386

occupied the niche that is suitable for D. r. bugensis. In Lake Balaton, D. polymorpha was 387

also found to rapidly colonize new substrata before the appearance of D. r. bugensis (Balogh 388

et al., 2008), meaning that primarily the colonization speed is due to substratum saturation 389

and not dependent on the type of dreissenid species. Nevertheless, the presence of competitors 390

can significantly influence the process.

391

During the colonization process, the D. polymorpha population was mainly confined to 392

a mid-size range (8-14 mm) that grew slowly. On the spat side, the D. polymorpha population 393

could not be renewed, since after one year of colonization, D. polymorpha settlers were rarely 394

found. On the adult side, the higher density of large (> 20 mm) D. r. bugensis individuals 395

suggested more sexually mature D. r. bugensis in the sample. Pressure from both sides might 396

lead to a decline, if not collapse, of the population of D. polymorpha during the simultaneous 397

(17)

17 colonization of the two dreissenid species. Before D. r. bugensis was introduced to the lake, 398

D. polymorpha grew larger (up to 2.4 cm, Balogh et al., 2008), which, considering that the 399

general conditions of the lake have not changed in the past several decades, suggests that D. r.

400

bugensis negatively affects D. polymorpha population development. The possible influence of 401

the D. r. bugensis population on D. polymorpha growth may also be strengthened by the fact 402

that neither in our sampling points nor at other sites of the lake (Balogh C., unpublished 403

observation) can D. polymorpha individuals >2 cm be found.

404

From 2011, large D. r. bugensis individuals (between 20 and 28 mm), showing 405

heterogeneous frequency, were counted on experimental substrata. These animals must have 406

arisen from a compact group in which sizes were between 12 and 19 mm in 2010. The 407

reduction of the number of these large animals in 2012 could be due to the decline of the first 408

generation after the third year of settling, since, as found earlier in Lake Balaton (Balogh et 409

al., 2008) and other water bodies (Whitney et al., 1996), the lifetime of dreissenids often does 410

not exceed 3–4 years. However, different access to resources (food, oxygen), due to different 411

site positions of the individuals in the substratum (i.e. at the bottom of the multilayer 412

aggregation, or on that substratum side, which is relatively hidden from the water current), 413

may contribute to the heterogeneous size distribution observed in large animals.

414

By the end of the study (August, 2012), the average length of the two species had 415

equalized, which may be explained, on the one hand, by the decreasing number of large size 416

(> 20 mm) individuals and the increasing number of young settling (3-10 mm) D. r. bugensis 417

individuals, and on the other hand, by the smoothly growing population of the mid-size (10- 418

17 mm) first colonizer generation of D. polymorpha.

419

While the veliger density was not significantly different in the summers of 2010–2012, 420

the density of new settlers declined, suggesting that the progression of surface occupation, 421

and hence, increasing saturation by growing dreissenids did not favor new settling 422

(18)

18 generations. Parallel to the decrease in free settling places, the growing shell surfaces of 423

earlier colonized individuals provided novel surfaces for larvae to attach, as observed from 424

2011. According to the experimental study of Tošenovský and Kobak (2016), the lower initial 425

distances between settled mussels offered a higher possibility for aggregation, suggesting that 426

as habitats are narrowed during the D. r. bugensis invasion, multilayer aggregation is 427

facilitated. Substratum saturation by dreissenids was calculated to be completed around two 428

years after substrata implanted to the lake, after which the competition for surfaces becomes 429

more intense between the two dreissenids. This was confirmed recently in an experiment 430

(Dzierzynska-Bialonczyk et al., 2017), where the formation of D. polymorpha aggregations 431

was found as a consequence of the lack of available alternative attachment sites. On the other 432

hand, saturation progression and in contrast, uncovered sites on experimental substratum 433

found until the last year of the study (four years after substratum implantation), suggested that 434

the shell of dreissenids appeared as an alternative attachment surface that promotes multilayer 435

aggregation.

436 437

Abiotic and biotic factors that would explain the success of D. r. bugensis colonization over 438

D. polymorpha 439

440

As the temperature dropped in the late autumn of 2009, the density of veligers 441

decreased, and subsequently, the colonization of post-veligers was reduced. From 2010 442

onward, the new settlers were almost exclusively D. r. bugensis, which might be because D. r.

443

bugensis starts spawning earlier at lower temperatures (4–9°C, Claxton and Mackie, 1998;

444

Roe and MacIsaac 1997; Stoeckmann, 2003; Nalepa et al., 2010) than D. polymorpha (above 445

9°C, Sprung, 1987). In Lake Balaton, veligers could be found in the water column at 7.3 °C 446

(late November), but they were missing in December when the temperature dropped below 7 447

(19)

19

°C. Interestingly, before the D. r. bugensis invasion, veligers were usually missing in the cold 448

season, even in October, when the water temperature was higher than 8 °C (Balogh et al., 449

2008). The water temperature may explain the inability of D. polymorpha to produce larvae 450

between October and April in Lake Balaton, and could be an environmental factor providing 451

D. r. bugensis with an advantage for earlier spawning, and thus a settling opportunity.

452

From 2011 to 2012, the abundance of the two dreissenids on experimental substrata 453

remained unchanged. This can be explained by the weather extremities (high water 454

temperature, water level fluctuation, waves), which are less endurable for D. r. bugensis 455

(Karatayev et al., 2013b), partly due to the mild attachment strength and fragility of the shell 456

(Peyer et al., 2009; Casper and Johnson, 2010), and their low tolerance to temperatures above 457

30.5 °C (Spidle et al., 1995; Thorp et al., 1998; Karatayev et al., 1998). Conversely, in 2013, 458

the lake was spared from weather extremities. In spring, the water level became so high that 459

the local substrata suitable for colonization were submerged again. This allowed for the 460

ongoing domination of D. r. bugensis on the rocks of the riprap in our sampling sites, 461

resulting in almost total displacement of D. polymorpha.

462

Since there is no evidence of selective erasure from the substrata, or early death of D.

463

polymorpha the different size distribution found between the two dreissenids after one year of 464

substratum deployment could be due to the faster growth of D. r. bugensis. A similar 465

difference in the growth rate between the two dreissenids was reported from the Laurentian 466

Great Lakes (Jarvis et al., 2000; Diggins, 2001; Stoeckmann, 2003), which was attributed to 467

the lower respiration and higher filtration rate of D. r. bugensis. A lower respiration rate 468

enables D. r. bugensis to reduce the energetic expenditure on maintenance, and therefore 469

promotes faster growth and ensures better chances for survival (Baldwin et al., 2002;

470

Stoeckmann, 2003). D. r. bugensis grows faster and is heavier at the same shell length, so 471

generally has a larger shell length and body mass than D. polymorpha (Mills et al. 1996, 472

(20)

20 Jarvis et al. 2000, Diggins, 2001; Stoeckmann 2003; Karatayev et al., 2010b). It is known 473

from field (Karatayev et al., 1998; Stoeckmann, 2003; Orlova et al., 2005), and from 474

experimental studies (Stoeckmann and Garton, 2001; Baldwin et al. 2002) that the growth and 475

body mass of D. polymorpha declines more on a poor-quality diet and in the presence of high 476

suspended material concentration, than that of D. r. bugensis. Our experimental and local 477

sampling points are situated in the oligotrophic part of the lake, and are characterized by low 478

food (chlorophyll-a concentration was 5.4±3.15 μg L-1), and high suspended material 479

concentrations (25-600 mg dry weight L-1, G.-Tóth et al., 2011), which may explain the faster 480

growth of D. r. bugensis.

481

There is no evidence that D. polymorpha reaches sexual maturation at lower sizes than 482

D. r. bugensis, but if this is the case, this could also explain why it develops slowly thereafter.

483

In the years after substratum deployment, however, among the second and subsequent 484

generations of post-veligers attached on the experimental substrata, D. polymorpha 485

individuals were rarely found, suggesting that the aforementioned assumption might be false.

486

Instead, the recruitment of D. r. bugensis implies that more larvae, and a larger number of 487

individuals producing larvae belonging to this species, were present in the surroundings from 488

2010. Since local circumstances that would have influenced the selective depletion of D.

489

polymorpha larvae are not known, the declined colonization of this species may due be to the 490

decreasing number of sexually mature D. polymorpha individuals.

491

Fish and bird predation, which concerns mainly medium sized (8–17 mm) dreissenids 492

(Czarnołęski, et al., 2006), regulates the dreissenid population in Lake Balaton (Ponyi, 1985;

493

Specziár et al., 1997, Balogh et al., 2008). This, besides natural death, might cause the 494

heterogeneous size distribution of the larger/older (> 20 mm) D. r. bugensis individuals found 495

on experimental substrata. However, we do not know whether predation evokes any species 496

selectivity. Other factors such as ice scours (MacIsaac, 1996; Chase and Bailey, 1999; Balogh 497

(21)

21 et. al, 2008), and parasites (Molloy et al., 1997), considered as potential regulators of 498

dreissenid population dynamics (Strayer and Malcom, 2006), are less feasibly involved in our 499

study.

500 501

Comparison of dreissenid invasion dynamics in the eastern basin of Lake Balaton with other 502

lakes 503

504

The colonization dynamics of dreissenids and the replacement of D. polymorpha with 505

D. r. bugensis happened similarly in our experimental and local study sites to that found in the 506

entire eastern basin of Lake Balaton (Balogh C., unpublished), where oligo-mesotrophic 507

conditions are uniform (chlorophyll-a concentration: 2-3 μg L-1, Sebestyén et al., 2017), and 508

also in other European shallow reservoirs (Orlova et al., 2004; Heiler et al., 2013), and lakes 509

(Matthews et al., 2014; Bij de Vaate et al., 2013). In these water bodies, D. r. bugensis almost 510

entirely replaced D. polymorpha 3–4 years after its appearance. In the Laurentian Great 511

Lakes, the progression of D. r. bugensis, mainly in deep areas, was much slower: it took more 512

than 10 years, but finally led to a significant reduction of D. polymorpha population 513

(Patterson et al., 2005; Watkins et al., 2007; Dermott and Dow, 2008; Nalepa et al. 2010). By 514

contrast, in shallow areas of the Lakes, like the western basin of Lake Erie, the two 515

dreissenids have lived together for a long time (Karatayev et al., 2014). Recently, we also 516

found the co-existence of the two species in the western basin of Lake Balaton (Balogh C., 517

unpublished), where the water trophity (chlorophyll-a concentration: 5-7 μg L-1, Sebestyén et 518

al., 2017) similar to that of the western basin of Lake Erie (Barbiero and Tuchman, 2004), is 519

more eutrophic than the Eastern basin. Hence, in shallow lakes, it seems that rapid 520

replacement of D. polymorpha with D. r. bugensis more likely happens if food availability is 521

limited.

522

(22)

22 The difference between the colonization history of dreissenids in North America and in 523

Europe is that North America was invaded by the two species simultaneously (Carlton, 2008, 524

Mills et al., 1993), whereas in Europe, D. polymorpha colonized and became the dominant 525

macroinvertebrate in the benthic community well before the appearance of D. r. bugensis 526

(Van der Velde et al., 2010). Therefore, another possible explanation for the different duration 527

of the replacement found between the continents is that parallel invasion could evoke a longer 528

struggle for place and resources between dreissenids having similar ecological requirements.

529

Where D. polymorpha was the first and prevailing dreissenid for years, the algal biomass and 530

hence the trophic state of the habitat reduced, a condition that is much more tolerable for D. r.

531

bugensis than D. polymorpha. Therefore, if D. r. bugensis is introduced to a habitat where D.

532

polymorpha has already been colonized for a long time, the new invader has a competitive 533

advantage as it better tolerates poor food conditions (Karatayev et al., 1998; Baldwin et al., 534

2002; Stoeckmann, 2003; Orlova et al., 2005). Similarly, the invasion of D. polymorpha and 535

its competitor, the amphipod Chelicorophium curvispinum (Sars, 1895) gave different results 536

in Lake Balaton and the river Rhine. In Lake Balaton, where the two species were introduced 537

together (Sebestyén, 1938), they have lived side by side for a long time (Balogh et al., 2008).

538

By contrast, in the river Rhine, the much later introduced amphipod gradually suppressed the 539

mussel population over several years (van den Brink et al., 1993; van der Velde et al., 1994).

540

Hence, it is possible that the progression and fate of the D. r. bugensis invasion highly 541

depends on whether the invader comes simultaneously with or later than its competitors. In 542

summary, it can be predicted that if D. r. bugensis appears later than D. polymorpha,in a 543

shallow lake where food availability is low, then the replacement becomes rapid.

544 545 546

Conclusion 547

(23)

23 548

The colonization process of dreissenids (D. polymorpha and D. r. bugensis) on 549

experimental and local substrata, simultaneously, at the time of the D. r. bugensis invasion, 550

revealed that the new invasive species was very successful against its congener, the formerly 551

dominating D. polymorpha in a large European shallow lake, where environmental conditions 552

favor the settlement of the new invader. The differences found in the speed of replacement 553

process between habitats in Europe and the Laurentian Great Lakes raise the necessity of 554

running a cross-system analysis involving many lakes that have dreissenid population data.

555

This would support the relevance of our hypothesis that the habitat previously occupied and 556

modified by D. polymorpha facilitates the conduction of rapid invasion by D. r. bugensis.

557

Detailed population analysis revealed that the success of D. r. bugensis is due to the 558

increasing number of large (> 20 mm) reproducing individuals and the consequently recruited 559

generations. The introduction of new substrata (e.g. setting piers, ship and boat stations) more 560

likely favors the progression of D. r. bugensis invasion, which in turn implies that the proper 561

selection of substratum type, or coating them with material inhibiting dreissenid attachment, 562

might contribute to reducing or delaying the propagation of D. r. bugensis in newly invaded 563

habitats. Nevertheless, it is necessary to study the reason for competition (ability to 564

predominate, evidence of the impact of environmental factors, such as food availability and 565

combined abiotic status) in the future, so as to make predictions about the invasion of 566

dreissenids into shallow lakes.

567 568

Acknowledgements 569

570

Authors are indebted to Mrs. Henrietta Szabó, Mrs. Tünde Klein-Polgárdi, Ms. Judit 571

Nédli and Ms. Szandra Purgel for their excellent technical assistance in the sample processing 572

(24)

24 and to Mr. Géza Dobos and Mr. Péter Harmati for their generous help in sampling. The text 573

was language edited by the Proof-Reading-Service.com Ltd, Letchworth Garden City, 574

Hertfordshire, UK. Study was financially supported by the TÁMOP-4.2.2.A-11/1/KONV- 575

2012-0038, the GINOP-2.3.2-15-2016-00019, the Balaton monitoring framework of the 576

Hungarian Academy of Sciences, and the bilateral project of MTA (NKM-38/2014). Long 577

term temperature dataset was obtained from LIFE08 ENV/IT/000339.

578 579

References 580

581

Ackerman, J.D., Smith, S.J., Nichols, S.J., Claudi, R., 1994. A review of the early life history 582

of zebra mussels (Dreissena polymorpha): comparisons with marine bivalves. Canadian 583

Journal of Zoology 72, 1169–1179.

584

Aldridge, D.C., Elliott, P., Moggridge, G.D., 2004. The recent and rapid spread of the zebra 585

mussel (Dreissena polymorpha) in Great Britain. Biological Conservation 119, 253- 586

261.

587

Aldridge, D.C., Samantha, H., Froufe, E., 2014. The Ponto-Caspian quagga mussel, 588

Dreissena rostriformis bugensis (Andrusov, 1897), invades Great Britain. Aquatic 589

Invasions 9, 529–535.

590

Baldwin, B.S., Mayer, M.S., Dayton, J., 2002. Comparative growth and feeding in zebra and 591

quagga mussels (Dreissena polymorpha and Dreissena bugensis): Implications for 592

North American lakes. Canadian Journal of Fisheries and Aquatic Sciences 59, 680–

593

694.

594

Balogh, C., G.-Tóth, L., 2009. A Balaton bevonatlakó gerinctelen állatvilágának vizsgálata a 595

2008. évben. In Bíró, P., J. A. Banczerowsky (eds), Balaton kutatásának 2008. évi 596

eredményei. MTA, Budapest: 45-53 pp (In Hungarian).

597

(25)

25 Balogh, C., Muskó, I.B., G.-Tóth, L., Nagy, L., 2008. Quantitative trends of zebra mussels in 598

Lake Balaton (Hungary) in 2003-2005 at different water levels. Hydrobiologia 613, 57- 599

69.

600

Barbiero, R.P., Tuchman,M.L., 2004. Long-term dreissenid impacts on water clarity in Lake 601

Erie. Journal of Great Lakes Research 30, 557–565.

602

Bij de Vaate, A., 1991. Distribution and aspects of population dynamics of the zebra mussel, 603

Dreissena polymorpha (Pallas, 1771), in the Lake Ijsselmeer area (The Netherlands).

604

Oecologia 86, 40-50.

605

Bij de Vaate, A., 2010. Some evidence for ballast water transport being the vector of the 606

quagga mussel (Dreissena rostriformis bugensis Andrusov 1897) introduction into 607

Western Europe and subsequent upstream dispersal in the river Rhine. Aquatic 608

Invasions 5, 207–209.

609

Bij de Vaate, A., van der Velde, G., Leuven, S. E.W.R., Heiler, K.C.M., 2013. Spread of the 610

Quagga Mussel (Dreissena rostriformis bugensis) in Western Europe In Nalepa, T. F., 611

D. Schloesser (eds), Quagga and Zebra Mussels: Biology, Impacts, and Control. CRC 612

Press, 83-92 pp.

613

Carlton J.T., 2008. The zebra mussel Dreissena polymorpha found in North America in 1986 614

and 1987. Journal of Great Lakes Research 34, 770–773.

615

Carpenter, S.R., Kitchell, J.F., 1993. The Trophic Cascade in Lakes. Cambridge University 616

Press, New York.

617

Casper, A.F., Johnson, L.E., 2010. Contrasting shell/tissue characteristics of Dreissena 618

polymorpha and Dreissena bugensis in relation to environmental heterogeneity in the 619

St. Lawrence River. Journal of Great Lakes Research 36, 184–189.

620

(26)

26 Chase, M.E., Bailey R.C., 1999. The ecology of the zebra mussel (Dreissena polymorpha) in 621

the lower Great Lakes of North America: I. Population dynamics and growth. Journal of 622

Great Lakes Research 25, 107–121.

623

Cianfanelli, S., Lori, E., Bodon, M., 2010. Dreissena polymorpha: current status of 624

knowledge about the distribution in Italy. In: G. van der Velde, S. Rajagopal, A. Bij de 625

Vaate (Eds), The Zebra Mussel in Europe, Backhuys/Margraf Publishers, 93-100.

626

Claudi, R, Mackie, G.L., 1994. Practical manual for zebra mussel monitoring and control.

627

Lewis Publishers, CRC Press, 227 pp.

628

Claxton, W.T., Mackie, G.L., 1998. Seasonal and depth variation in gametogenesis and 629

spawning of Dreissena polymorpha and Dreissena bugensis in eastern Lake Erie.

630

Canadian Journal of Zoology 76, 2010–2019.

631

Collas, F.P.L., Karatayev A.Y., Burlakova L.E., Leuven R.S.E.W., 2016. Detachment rates of 632

dreissenid mussels after boat hull-mediated overland dispersal. Hydrobiologia, doi:

633

10.1007/s10750-016-3072-4.

634

Crooks, J.A., 1998. Habitat alteration and community-level effects of an exotic mussel, 635

Musculista senhousia. Marine Ecology Progress Series 162, 137–152.

636

Czarnołęski, M., Kozłowski, J., Kubajak, P., Lewandowski, K., Müller, T., Stańczykowska, 637

A., Surówka, K., 2006: Crosshabitat differences in crush resistance and growth pattern 638

of zebra mussels (Dreissena polymorpha): effects of calcium availability and predator 639

pressure. Archiv für Hydrobiologie 165, 191–208.

640

DAISIE, 2003. Delivering Alien Invasive Species In Europe, funded by the European 641

Commission under the Sixth Framework Programme, Contract Number: SSPI-CT- 642

2003-511202.

643

De Gelder, S., Van der Velde, G., Platvoet, D., Leung, N., Dorenbosch, M., Hendriks, 644

H.W.M., Leuven, R.S.E.W., 2016. Competition for shelter sites: Testing a possible 645

(27)

27 mechanism for gammarid species displacements, Basic and Applied Ecology 17, 455- 646

462.

647

Dermott, R., Dow, J., 2008. Changing benthic fauna of Lake Erie between 1993 and 1998. In 648

Checking the Pulse of Lake Erie. In Munawar, M., R. Heath (eds), Goodwords Books.

649

New Delhi, India: 409–438 pp.

650

Dick, J.T.A., Platvoet, D., 2000. Invading predatory crustacean Dikerogammarus villosus 651

eliminates both native and exotic species Proceedings of the Royal Society London B 652

267, 977-983.

653

Diggins, T.P., 2001. A seasonal comparison of suspended sediment filtration by quagga 654

(Dreissena bugensis) and zebra (D. polymorpha) mussels. Journal of Great Lakes 655

Research 27, 457–466.

656

Dzierzynska-Bialonczyk, A., Skrzypczak, A., Kobak, J., 2017. Happy together? Avoidance of 657

conspecifics by gregarious mussels. Current Zoology, doi: 10.1093/cz/zox022 658

Gallardo, B., Aldridge, D.C., 2015. Is Great Britain heading for a Ponto-Caspian Invasional 659

Meltdown? Journal of Applied Ecology 52, 41-49.

660

G.-Tóth, L., Parpala, L., Balogh, C., Tátrai, I., Baranyai, E., 2011. Zooplankton community 661

response to enhanced turbulence generated by water-level decrease in Lake Balaton, the 662

largest shallow lake in Central Europe. Limnology and Oceanography 56, 2211-2222.

663

Gutiérrez, J.L., Jones, C.G., Strayer, D.L., Iribarne, O.O., 2003. Mollusks as ecosystem 664

engineers: the role of shell production in aquatic habitats. Oikos 101, 79–90.

665

Heiler, K.C.M., Brandt, S., Albrecht, C., Hauffe, T., Wilke, T., 2012. A new approach for 666

dating introduction events of the quagga mussel (Dreissena rostriformis bugensis).

667

Biological Invasions 14, 1311–1316 668

(28)

28 Heiler, K.C.M., Bij de Vaate, A., Ekschmitt, K., von Oheimb, P.V., Albrecht, C., Wilke, T., 669

2013. Reconstruction of the early invasion history of the quagga mussel (Dreissena 670

rostriformis bugensis) in Western Europe. Aquatic Invasions 8, 53–57 671

Jarvis, P., Dow, J., Dermott, R., Bonnell, R., 2000. Zebra (Dreissena polymorpha) and quagga 672

mussel (Dreissena bugensis) distribution and density in Lake Erie 1992–1998. Canadian 673

Technical Report of Fisheries and Aquatic Sciences 2304, 1-46.

674

Karatayev, A.Y., Burlakova, L.E., Padilla, D.K., 1998. Physical factors that limit the 675

distribution and abundance of Dreissena polymorpha (Pall.). Journal of Shellfish 676

Research 17, 1219–1235.

677

Karatayev, A.Y., Burlakova, L.E., Padilla, D.K., 2010a. Dreissena polymorpha in Belarus:

678

history of spread, population biology, and ecosystem impacts. Chapter 9 in G. van der 679

Velde, S. Rajagopal and A. Bij de Vaate (eds), The Zebra Mussel in Europe. Backhuys 680

Publishers, Leiden/ Margraf Publishers, Weikersheim. p. 101-112.

681

Karatayev, A.Y., Mastitsky, S.E., Padilla, D.K., Burlakova, L.E., Hajduk, M.M., 2010b.

682

Differences in growth and survivorship of zebra and quagga mussels: size matters.

683

Hydrobiologia. 668, 183-194.

684

Karatayev, A.Y., Burlakova, L.E., Mastitsky, S.E., Padilla, D.K., Mills, E.L., 2011.

685

Contrasting rates of spread of two congeners, Dreissena polymorpha and Dreissena 686

rostriformis bugensis at different spatial scales. Journal of Shellfish Research 30, 923–

687

931.

688

Karatayev, V.A., Karatayev A.Y., Burlakova, D.K.P., 2013a. Lakewide dominance does not 689

predict the potential for spread of dreissenids. Journal of Great Lakes Research 39, 622–

690

629.

691

Karatayev, A.Y., Burlakova, L.E., Padilla, D.K., 2013b. General overview of zebra and 692

quagga mussels: what we know and do not know. In Nalepa, T. F., D. Schloesser (eds), 693

(29)

29 Quagga and Zebra Mussels: Biology, Impacts, and Control, 2nd Edition. CRC Press, 694

Boca Raton: FL, 695–703 pp.

695

Karatayev, A.Y., Burlakova, L.E., Pennuto, C., Ciborowski, J., Karatayev, V.A., Juette, P., 696

Clapsadl, M., 2014. Twenty five years of changes in Dreissena spp. populations in Lake 697

Erie. Journal of Great Lakes Research 40, 550-559.

698

Kirpichenko, M.Y., 1964. Phenology, population dynamics, and growth of Dreissena larvae 699

in the Kuibyshev Reservoir. Tr. Inst. Biol. Vnutr. Vod Akad. Nauk SSSR 7(10),15-24 700

(in Russian).

701

MacIsaac, H.J., 1996. Population structure of an introduced species (Dreissena polymorpha) 702

along a waveswept disturbance gradient. Oecologia 105, 484–492.

703

Madon, S.P., Schneider, D.W., Stoeckel, J.A., Sparks, R.E., 1998. Effects of inorganic 704

sediment and food concentrations on energetic processes of the zebra mussel, Dreissena 705

polymorpha: implication for growth in turbid rivers. Canadian Journal of Fisheries and 706

Aquatic Sciences 55, 401-413.

707

Majoros, G., 2009. Invazív kagylófajok terjeszkedése a Balatonban: esetismertetés és a 708

probléma felvetése. Halászatfejlesztés - Fisheries and Aquaculture Development 32,57- 709

64 (In Hungarian).

710

Marsden, J.E., 1992. Standard protocols for monitoring and sampling zebra mussels. Illinois 711

Natural History Survey Biological Notes 138, 1-40.

712

Matthews, J. , Van der Velde, G., Bij de Vaate, A., Collas, F.P.L., Koopman K.R., Leuven, 713

R.S.E.W., 2014. Rapid range expansion of the invasive quagga mussel in relation to 714

zebra mussel presence in The Netherlands and Western Europe. Biological Invasions 715

16, 23-42.

716

Mills, E.L., Dermott, R.M., Roseman, E.F., Dustin, D., Mellina, E., Conn, D.B., Spidle, A.P., 717

1993. Colonization, ecology, and population structure of the ‘‘quagga’’ mussel 718

(30)

30 (Bivalvia: Dreissenidae) in the lower Great Lakes. Canadian Journal of Fisheries and 719

Aquatic Sciences 50, 2305–2314.

720

Mills, E.L., Rosenberg, G., Spidle, A.P., Ludyanskiy, M., Pligin, Y. May, B., 1996. A review 721

of the biology and ecology of the quagga mussel (Dreissena bugensis), a second species 722

of freshwater Dreissenid introduced to North America. American Zoologist 36, 271–

723

286.

724

Molloy, D. P., Karatayev, A.Y., Burlakova, L.E., Kurandina, D.P., Laruelle, F., 1997.

725

Natural enemies of zebra mussels: predators, parasites, and ecological competitors.

726

Reviews in Fishery Science 5, 27–97.

727

Nalepa, T.F., Fanslow, D.L., Pothoven, S.A., 2010. Recent changes in density, biomass, 728

recruitment, size structure, and nutritional state of Dreissena populations in southern 729

Lake Michigan. Journal of Great Lakes Research 36, 5–19.

730

Orlova, M.I., 2013. Origin and Spread of Quagga Mussels (Dreissena rostriformis bugensis) 731

in Eastern Europe with Notes on Size Structure of Populations Pp. 93-102 In: Quagga 732

and zebra Mussels: Biology, Impacts, and Control (T.F. Nalepa, D. Schloesser, eds.), 733

CRC Press.

734

Orlova, M.I., Muirhead, J.R., Antonov, P.I., 2004. Range expansion of quagga mussels 735

Dreissena rostriformis bugensis in the Volga River and Caspian Sea basin. Aquatic 736

Ecology 38, 561–573.

737

Orlova, M.I., Therriault, T.W., Antonov, P.I., Shcherbina, G.K., 2005. Invasion ecology of 738

quagga mussels (Dreissena rostriformis bugensis): A review of evolutionary and 739

phylogenetic impacts. Aquatic Ecology 39, 401–418.

740

OTA, (Office of Technology Assessment), 1993. Harmful non-indigenous species in the 741

United States. Publication No. OTA-F-565, US Government Printing Office, 742

Washington D. C.

743

(31)

31 Patterson, M.W.R., Ciborowski, J.J.H., Barton, D.R., 2005. The distribution and abundance of 744

Dreissena species (Dreissenidae) in Lake Erie, 2002. Journal of Great Lakes Research 745

31,223-237.

746

Palau Ibars, A., Cia Abaurre, I., Casas Mulet, R., Rosico Ramón, E., 2010. Zebra mussel 747

distribution and habitat preference at the Ebro basin (north east Spain), Chap. 10. In 748

Van der Velde, G., S. Rajagopal, A. Bij de Vaate (eds), The Zebra Mussel in Europe.

749

Backhuys Publishers, Leiden/Margraf Publishers, Weikersheim: 113–118.

750

Pimentel, D., Zuniga, R., Morrison, D., 2005. Update on the environmental and economic 751

costs associated with alien-invasive species in the United States. Ecological Economics 752

52, 273-288.

753

Petrie, S.A., Knapton, R.W., 1999. Rapid increase and subsequent decline of zebra and 754

quagga mussels in Long Point Bay, Lake Erie: possible influence of waterfowl 755

predation. Journal of Great Lakes Research 25, 772–782.

756

Peyer, S.M., McCarthy, A.J., Lee, C. E., 2009. Zebra mussels anchor byssal threads faster and 757

tighter than quagga mussels in flow. Journal of Experimental Biology 212, 2027–2036.

758

Platvoet, D., Dick, J.T.A., MacNeil, C., van Riel, M.C., van der Velde, G., 2009. Invader- 759

invader interactions in relation to environmental heterogeneity leads to zonation of two 760

invasive amphipods, Dikerogammarus villosus (Sowinsky) and Gammarus tigrinus 761

Sexton: amphipod pilot species project (AMPIS) report 6. Biological Invasions 11, 762

2085–2093.

763

Pollux, B.J.A., Minchin, D., Van der Velde, G., Van Alen, T., Moon-Van der Staay, S., 764

Hackstein, J.H.P., 2003. Zebra mussels (Dreissena polymorpha) in Ireland, AFLP 765

fingerprinting and boat traffic both indicate an origin from Britain. Freshwater Biology 766

48, 1127-1139.

767

Ábra

Fig. 1. Site map of the study site at Lake Balaton and the Tihany peninsula. Location where 860
Fig.  3.  Density  of  dreissenid  veliger  larvae  in  water,  and  recruitment  dynamics  of  post-post-876
Fig.  5.  Relative  abundance  of  D.  r.  bugensis  in  the  dreissenid  population  on  experimental 900
Fig.  7. Shell length  of dreissenids  (size &gt;2 mm) on  experimental  substrata. Each data point 914
+2

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

protected plant species occurred in 69% of graveyards that hosted Spiraea crenata.. Substantial differences could be observed in frequency,

Both methods pointed out that (1) the high proportions of deciduous (mainly beech and oak) trees on the sampling units were preferred by the majority of fungal species, (2)

While the colonial and moderately attached STGs correlated positively to nitrate (Fig. 2), the S3, S4 and S5 sized, slow moving, colonial and weakly attached CTGs showed positive

(2018) Within-generation and transgenerational plasticity in growth and regeneration of a subordinate annual grass in a rainfall experiment.. 2-4, H-2163

47 Table 1 Multiple-site Jaccard dissimilarity (multiple-D J ; total beta diversity), species replacement (multiple-Repl PJ ) and richness difference (multiple-Rich PJ ) sensu

data of four common fish species (bleak, roach, perch, pumpkinseed sunfish) collected from 34.. three closely related sites were

First DCA axis scores (SD units) and significant assemblage zones of the diatom (GAL-d 1-7) and chironomid (GAL-ch 1-5) records plotted together with selected explanatory

Multiple Potential Natural Vegetation Model (MPNV), a novel approach supported 30.. restoration prioritization satisfying both ecological (sustainability and nature